Jump to content

Series (mathematics)

From Wikipedia, the free encyclopedia

This is an old revision of this page, as edited by Hesselp (talk | contribs) at 21:24, 28 April 2017 (Reverter D.Lazard, please write in Talk: the arguments presented by each of that '4 editors'.). The present address (URL) is a permanent link to this revision, which may differ significantly from the current revision.

In mathematics (calculus), the word series is primarily used as adjective specifying a certain kind of expressions denoting numbers (or functions).
Symbolic forms like        and     oder    expressing a number as the limit of the partial sums of sequence , are called series expression.  'Series expression' is often shortened to just 'series'.

Secondly, series is used as an adjective in series representation, denoting the kind of representation (of a number or a function) as a limit of the partial sums of a given sequence.

Thirdly, series is used, again as an adjective, in series expansion. Being a special type of series representation (of functions, not numbers).   For instance:
the Maclaurin expansion of a given function  and  the Fourier expansion of a given function  are series expansions.

Finally, (the noun) series can be synonymous with sequence.  Cauchy defined the word series by "an infinite sequence of real numbers". [1]   The use of the word 'series' for 'sequence' has a long tradition, with analogons in other languages, but seems to be considered as somewhat outdated.

The rather widespread idea about the existence of a mathematical notion (a definable mathematical entity, called 'series'), 'associated'  in some way with a given number sequence, with its partial sums sequence, and with the eventual limit thereof,   misses a sound base. [2] [3]

The study of the series representation is a major part of mathematical analysis. With this tool, irrationals can be described/defined by means of (the limit of) a relatively easy descriptable sequence of rationals.  
This kind of representation is used in most areas of mathematics, even for studying finite structures (such as in combinatorics), through generating functions. In addition to their ubiquity in mathematics, the series representation is also widely used in other quantitative disciplines such as physics, computer science, statistics and finance.

Situations in which the word  'series'  is used

Definitions, common wordings

Given a infinite sequence  with terms et cetera (or starting with ) for which addition is defined, the sequence
    is called  the sequence of partial sums of sequence .
Alternative notation:      .   Alternative name: the sum sequence of (sequence) [4].
Example: The sequence (1, 2, 3, 4, ···)  is the sum sequence of (1, 1, 1, 1, ··· ); being the sum sequence of (1, 0, 0, 0, ··· );   this can be extended in both directions.

A series, short for series expression, is a written expression using mathematical signs, consisting of
- an expression denoting the function that maps a given sequence on the limit of its sum sequence,  combined with
- an expression denoting an infinite sequence (with addition and distance defined).
Examples:       (plusses-bullets notation),         (capital-sigma notation).
Sometimes, the same symbolic forms are used to denote the sum sequence of  , instead of the value of its eventual limit.

A sequence with a converging sum sequence is called summable.  The finite limit is called sum of the sequence.

A valid series expression has a summable sequence as its argument (and denotes a value). Otherwise the expression is void. Traditional wordings are: "convergent/divergent series expression" or "convergent/divergent series".

Convergent / divergent series   The combination convergent series shouldn't be interpreted literally, for an expression itself cannot be convergent or divergent.  By tradition  "Σ is a convergent series"  as well as  "series Σ converges"  are used to express that sequence is summable.   Similarly, "Σ is a divergent series"  and  "series Σ diverges"  are used to say that sequence is not summable.

Convergence test for series   Again, this traditional wording cannot be taken literally because 'series' is the name of an expression of a certain kind, not the name of a mathematical notion. An alternative is: summability test for sequences.

Absolute convergent series   This is the traditional naming for a sequence with summable absolute values of its terms. The alternative  absolute summable sequence  is not in common use.

Series Σ   and  sequence   are interchangeable in traditional clauses like:
- the sum of series Σ ,     the terms of series Σ ,     the (sequence of) partial sums of series Σ ,     the Cauchy product of series Σ and series Σ
- the series Σ is geometric, arithmetic, harmonic, alternating, non negative, increasing  (and more).

There is no standard interpretation for the limit of series Σ .

Series representation of numbers and functions

In some contexts the word 'series' shouldn't be seen as referring to a certain type of written symbolic expressions, but as referring to a special type of representation of numbers (and functions). Namely: defining a (irrational) number as the limit of the partial sums of a known infinite sequence of (rational or irrational) numbers.  And in the case of functions: defining a function as the limit of the partial sums of an infinite sequence of functions (which are seen as 'easier' or more elementary in one way or another than the function represented by the limit).
As comparable with the idea of series representation (or: infinite sum representation) can be seen:  the continued fraction representation  and  the infinite product representation (for numbers and functions).

R e d u c t i o n   o f   s u m s   a n d     p r o d u c t s
A sum of two numbers given in series representation,
a product of two numbers given in series representation, and
a product of two numbers, one of them given in series representation,
can be reduced according to:
     
          (sequence or sequence summable)
      .
The same applies for functions instead of numbers.

Series expansion of functions

The name 'series expansion' is used for a special type of series representation of functions. (Not applicable to numbers.)
A series expansion is a series representation of a function, using a sequence of power functions of increasing degree, in one of its variables. Or functions like (for example) .
The labels Maclaurin series, Taylor series, Fourier series shouldn't be seen as denoting expressions but rather representations of the type series expansion. So Maclaurin series should be understood as Maclaurin expansion,  Fourier series as Fourier expansion, et cetera. [5]

Power series

The name power series can occur
- as synonym for Maclaurin expansion,  and
- denoting a series expression which includes an expression for a sequence of power functions with increasing degree.

Cauchy as a source of confusion

Cauchy, in his 'Cours d'Analyse' (1821) made an important, but quite subtile, distinction between the meaning of 'to converge' and 'being convergent':
- A sequence (French: suite) can converge to a limit.
- A sequence with converging partial sums, is called convergent by Cauchy (meaning 'summable')
Moreover, an infinite sequence with real numbers as terms, he called a series (French: série).
This imprudent choise caused permanent confusion around the use of the word 'series'  (e.g. in the German translations of 'Cours d'Analyse' of 1828 and 1885) [6]  until the present day.

Remark on the use of  'series'  and  'convergent / divergent'  in the sections below

Below, the words 'series' and 'convergent / divergent' are not always used conform the preceding descriptions. In such cases the context has to be taken into account to track down the intended meaning.


Examples

  • A geometric series is one where each successive term is produced by multiplying the previous term by a constant number (called the common ratio in this context). Example:
In general, the geometric series
converges if and only if .
The harmonic series is divergent.

and

converges if r > 1 and diverges for r ≤ 1, which can be shown with the integral criterion described below in convergence tests. As a function of r, the sum of this series is Riemann's zeta function.
converges if the sequence bn converges to a limit L as n goes to infinity. The value of the series is then b1L.
  • There are some elementary series whose convergence is not yet known/proven. For example, it is unknown whether the Flint Hills series
converges or not. The convergence depends on how well can be approximated with rational numbers (which is unknown as of yet). More specifically, the values of n with large numerical contributions to the sum are the numerators of the continued fraction convergents of , a sequence beginning with 1, 3, 22, 333, 355, 103993, ... (sequence A046947 in the OEIS). These are integers that are close to for some integer n, so that is close to 0 and its reciprocal is large. Max A. Alekseyev proved in 2011 that if the series converges then the irrationality measure of is lower than 2.5, which is much smaller than current known bound of 7.6063....[7][8]

Calculus and partial summation as an operation on sequences

Partial summation takes as input a sequence, ( an ), and gives as output another sequence, ( sn ). It is thus a unary operation on sequences. Further, this function is linear, and thus is a linear operator on the vector space of sequences, denoted Σ. The inverse operator is the finite difference operator, Δ. These behave as discrete analogs of integration and differentiation, only for sequences (functions of a natural number) instead of functions of a real variable. For example, the sequence (1, 1, 1, ...) has sequence (1, 2, 3, 4, ...) as its partial sums sequence, which is analogous to the fact that

In computer science it is known as prefix sum.

Properties of series

Series are classified not only by whether they converge or diverge, but also by the properties of the terms an (absolute or conditional convergence); type of convergence of the series (pointwise, uniform); the class of the term an (whether it is a real number, arithmetic progression, trigonometric function); etc.

Non-negative terms

When an is a non-negative real number for every n, the sequence SN of partial sums is non-decreasing. It follows that a series ∑an with non-negative terms converges if and only if the sequence SN of partial sums is bounded.

For example, the series

is convergent, because the inequality

and a telescopic sum argument implies that the partial sums are bounded by 2.

Absolute convergence

A series

is said to converge absolutely if the series of absolute values

converges. This is sufficient to guarantee not only that the original series converges to a limit, but also that any reordering of it converges to the same limit.

Conditional convergence

A series of real or complex numbers is said to be conditionally convergent (or semi-convergent) if it is convergent but not absolutely convergent. A famous example is the alternating series

which is convergent (and its sum is equal to ln 2), but the series formed by taking the absolute value of each term is the divergent harmonic series. The Riemann series theorem says that any conditionally convergent series can be reordered to make a divergent series, and moreover, if the an are real and S is any real number, that one can find a reordering so that the reordered series converges with sum equal to S.

Abel's test is an important tool for handling semi-convergent series. If a series has the form

where the partial sums BN = b0 + ··· + bn are bounded, λn has bounded variation, and lim λnBn exists:

then the series ∑ an is convergent. This applies to the pointwise convergence of many trigonometric series, as in

with 0 < x < 2π. Abel's method consists in writing bn+1 = Bn+1 − Bn, and in performing a transformation similar to integration by parts (called summation by parts), that relates the given series ∑ an to the absolutely convergent series

Convergence tests

  • n-th term test: If limn→∞ an ≠ 0 then the series diverges.
  • Comparison test 1 (see Direct comparison test): If ∑bn  is an absolutely convergent series such that |an | ≤ C |bn | for some number C  and for sufficiently large n , then ∑an  converges absolutely as well. If ∑|bn | diverges, and |an | ≥ |bn | for all sufficiently large n , then ∑an  also fails to converge absolutely (though it could still be conditionally convergent, e.g. if the an  alternate in sign).
  • Comparison test 2 (see Limit comparison test): If ∑bn  is an absolutely convergent series such that |an+1 /an | ≤ |bn+1 /bn | for sufficiently large n , then ∑an  converges absolutely as well. If ∑|bn | diverges, and |an+1 /an | ≥ |bn+1 /bn | for all sufficiently large n , then ∑an  also fails to converge absolutely (though it could still be conditionally convergent, e.g. if the an  alternate in sign).
  • Ratio test: If there exists a constant C < 1 such that |an+1/an|<C for all sufficiently large n, then ∑an converges absolutely. When the ratio is less than 1, but not less than a constant less than 1, convergence is possible but this test does not establish it.
  • Root test: If there exists a constant C < 1 such that |an|1/n ≤ C for all sufficiently large n, then ∑an converges absolutely.
  • Integral test: if ƒ(x) is a positive monotone decreasing function defined on the interval [1, ∞) with ƒ(n) = an for all n, then ∑an converges if and only if the integral  ∫1 ƒ(x) dx is finite.
  • Cauchy's condensation test: If an is non-negative and non-increasing, then the two series  ∑an  and  ∑2ka(2k) are of the same nature: both convergent, or both divergent.
  • Alternating series test: A series of the form ∑(−1)n an (with an ≥ 0) is called alternating. Such a series converges if the sequence an is monotone decreasing and converges to 0. The converse is in general not true.
  • For some specific types of series there are more specialized convergence tests, for instance for Fourier series there is the Dini test.

Series of functions

A series of real- or complex-valued functions

converges pointwise on a set E, if the series converges for each x in E as an ordinary series of real or complex numbers. Equivalently, the partial sums

converge to ƒ(x) as N → ∞ for each x ∈ E.

A stronger notion of convergence of a series of functions is called uniform convergence. The series converges uniformly if it converges pointwise to the function ƒ(x), and the error in approximating the limit by the Nth partial sum,

can be made minimal independently of x by choosing a sufficiently large N.

Uniform convergence is desirable for a series because many properties of the terms of the series are then retained by the limit. For example, if a series of continuous functions converges uniformly, then the limit function is also continuous. Similarly, if the ƒn are integrable on a closed and bounded interval I and converge uniformly, then the series is also integrable on I and can be integrated term-by-term. Tests for uniform convergence include the Weierstrass' M-test, Abel's uniform convergence test, Dini's test, and the Cauchy criterion.

More sophisticated types of convergence of a series of functions can also be defined. In measure theory, for instance, a series of functions converges almost everywhere if it converges pointwise except on a certain set of measure zero. Other modes of convergence depend on a different metric space structure on the space of functions under consideration. For instance, a series of functions converges in mean on a set E to a limit function ƒ provided

as N → ∞.

Power series

A power series is a series of the form

The Taylor series at a point c of a function is a power series that, in many cases, converges to the function in a neighborhood of c. For example, the series

is the Taylor series of at the origin and converges to it for every x.

Unless it converges only at x=c, such a series converges on a certain open disc of convergence centered at the point c in the complex plane, and may also converge at some of the points of the boundary of the disc. The radius of this disc is known as the radius of convergence, and can in principle be determined from the asymptotics of the coefficients an. The convergence is uniform on closed and bounded (that is, compact) subsets of the interior of the disc of convergence: to wit, it is uniformly convergent on compact sets.

Historically, mathematicians such as Leonhard Euler operated liberally with infinite series, even if they were not convergent. When calculus was put on a sound and correct foundation in the nineteenth century, rigorous proofs of the convergence of series were always required. However, the formal operation with non-convergent series has been retained in rings of formal power series which are studied in abstract algebra. Formal power series are also used in combinatorics to describe and study sequences that are otherwise difficult to handle; this is the method of generating functions.

Laurent series

Laurent series generalize power series by admitting terms into the series with negative as well as positive exponents. A Laurent series is thus any series of the form

If such a series converges, then in general it does so in an annulus rather than a disc, and possibly some boundary points. The series converges uniformly on compact subsets of the interior of the annulus of convergence.

Dirichlet series

A Dirichlet series is one of the form

where s is a complex number. For example, if all an are equal to 1, then the Dirichlet series is the Riemann zeta function

Like the zeta function, Dirichlet series in general play an important role in analytic number theory. Generally a Dirichlet series converges if the real part of s is greater than a number called the abscissa of convergence. In many cases, a Dirichlet series can be extended to an analytic function outside the domain of convergence by analytic continuation. For example, the Dirichlet series for the zeta function converges absolutely when Re s > 1, but the zeta function can be extended to a holomorphic function defined on   with a simple pole at 1.

This series can be directly generalized to general Dirichlet series.

Trigonometric series

A series of functions in which the terms are trigonometric functions is called a trigonometric series:

The most important example of a trigonometric series is the Fourier series of a function.

History of the theory of infinite series

Development of infinite series

Greek mathematician Archimedes produced the first known summation of an infinite series with a method that is still used in the area of calculus today. He used the method of exhaustion to calculate the area under the arc of a parabola with the summation of an infinite series, and gave a remarkably accurate approximation of π.[9][10]

Mathematicians from Kerala, India studied infinite series around 1350 CE.[11]

In the 17th century, James Gregory worked in the new decimal system on infinite series and published several Maclaurin series. In 1715, a general method for constructing the Taylor series for all functions for which they exist was provided by Brook Taylor. Leonhard Euler in the 18th century, developed the theory of hypergeometric series and q-series.

Convergence criteria

The investigation of the validity of infinite series is considered to begin with Gauss in the 19th century. Euler had already considered the hypergeometric series

on which Gauss published a memoir in 1812. It established simpler criteria of convergence, and the questions of remainders and the range of convergence.

Cauchy (1821) insisted on strict tests of convergence; he showed that if two series are convergent their product is not necessarily so, and with him begins the discovery of effective criteria. The terms convergence and divergence had been introduced long before by Gregory (1668). Leonhard Euler and Gauss had given various criteria, and Colin Maclaurin had anticipated some of Cauchy's discoveries. Cauchy advanced the theory of power series by his expansion of a complex function in such a form.

Abel (1826) in his memoir on the binomial series

corrected certain of Cauchy's conclusions, and gave a completely scientific summation of the series for complex values of and . He showed the necessity of considering the subject of continuity in questions of convergence.

Cauchy's methods led to special rather than general criteria, and the same may be said of Raabe (1832), who made the first elaborate investigation of the subject, of De Morgan (from 1842), whose logarithmic test DuBois-Reymond (1873) and Pringsheim (1889) have shown to fail within a certain region; of Bertrand (1842), Bonnet (1843), Malmsten (1846, 1847, the latter without integration); Stokes (1847), Paucker (1852), Chebyshev (1852), and Arndt (1853).

General criteria began with Kummer (1835), and have been studied by Eisenstein (1847), Weierstrass in his various contributions to the theory of functions, Dini (1867), DuBois-Reymond (1873), and many others. Pringsheim's memoirs (1889) present the most complete general theory.

Uniform convergence

The theory of uniform convergence was treated by Cauchy (1821), his limitations being pointed out by Abel, but the first to attack it successfully were Seidel and Stokes (1847–48). Cauchy took up the problem again (1853), acknowledging Abel's criticism, and reaching the same conclusions which Stokes had already found. Thomae used the doctrine (1866), but there was great delay in recognizing the importance of distinguishing between uniform and non-uniform convergence, in spite of the demands of the theory of functions.

Semi-convergence

A series is said to be semi-convergent (or conditionally convergent) if it is convergent but not absolutely convergent.

Semi-convergent series were studied by Poisson (1823), who also gave a general form for the remainder of the Maclaurin formula. The most important solution of the problem is due, however, to Jacobi (1834), who attacked the question of the remainder from a different standpoint and reached a different formula. This expression was also worked out, and another one given, by Malmsten (1847). Schlömilch (Zeitschrift, Vol.I, p. 192, 1856) also improved Jacobi's remainder, and showed the relation between the remainder and Bernoulli's function

Genocchi (1852) has further contributed to the theory.

Among the early writers was Wronski, whose "loi suprême" (1815) was hardly recognized until Cayley (1873) brought it into prominence.

Fourier series

Fourier series were being investigated as the result of physical considerations at the same time that Gauss, Abel, and Cauchy were working out the theory of infinite series. Series for the expansion of sines and cosines, of multiple arcs in powers of the sine and cosine of the arc had been treated by Jacob Bernoulli (1702) and his brother Johann Bernoulli (1701) and still earlier by Vieta. Euler and Lagrange simplified the subject, as did Poinsot, Schröter, Glaisher, and Kummer.

Fourier (1807) set for himself a different problem, to expand a given function of x in terms of the sines or cosines of multiples of x, a problem which he embodied in his Théorie analytique de la chaleur (1822). Euler had already given the formulas for determining the coefficients in the series; Fourier was the first to assert and attempt to prove the general theorem. Poisson (1820–23) also attacked the problem from a different standpoint. Fourier did not, however, settle the question of convergence of his series, a matter left for Cauchy (1826) to attempt and for Dirichlet (1829) to handle in a thoroughly scientific manner (see convergence of Fourier series). Dirichlet's treatment (Crelle, 1829), of trigonometric series was the subject of criticism and improvement by Riemann (1854), Heine, Lipschitz, Schläfli, and du Bois-Reymond. Among other prominent contributors to the theory of trigonometric and Fourier series were Dini, Hermite, Halphen, Krause, Byerly and Appell.

Generalizations

Asymptotic series

Asymptotic series, otherwise asymptotic expansions, are infinite series whose partial sums become good approximations in the limit of some point of the domain. In general they do not converge. But they are useful as sequences of approximations, each of which provides a value close to the desired answer for a finite number of terms. The difference is that an asymptotic series cannot be made to produce an answer as exact as desired, the way that convergent series can. In fact, after a certain number of terms, a typical asymptotic series reaches its best approximation; if more terms are included, most such series will produce worse answers.

Divergent series

Under many circumstances, it is desirable to assign a limit to a series which fails to converge in the usual sense. A summability method is such an assignment of a limit to a subset of the set of divergent series which properly extends the classical notion of convergence. Summability methods include Cesàro summation, (C,k) summation, Abel summation, and Borel summation, in increasing order of generality (and hence applicable to increasingly divergent series).

A variety of general results concerning possible summability methods are known. The Silverman–Toeplitz theorem characterizes matrix summability methods, which are methods for summing a divergent series by applying an infinite matrix to the vector of coefficients. The most general method for summing a divergent series is non-constructive, and concerns Banach limits.

Series in Banach spaces

The notion of series can be easily extended to the case of a Banach space. If xn is a sequence of elements of a Banach space X, then the series Σxn converges to x ∈ X if the sequence of partial sums of the series tends to x; to wit,

as N → ∞.

More generally, convergence of series can be defined in any abelian Hausdorff topological group. Specifically, in this case, Σxn converges to x if the sequence of partial sums converges to x.

Summations over arbitrary index sets

Definitions may be given for sums over an arbitrary index set I. There are two main differences with the usual notion of series: first, there is no specific order given on the set I; second, this set I may be uncountable.

If is a function from an index set I to a set G, then the "series" associated to is the formal sum of the elements over the index elements denoted by the

When the index set is the natural numbers , the function is a sequence denoted by . A series indexed on the natural numbers is an ordered formal sum and so we rewrite as in order to emphasize the ordering induced by the natural numbers. Thus, we obtain the common notation for a series indexed by the natural numbers

Families of non-negative numbers

When summing a family {ai}, i ∈ I, of non-negative numbers, one may define

When the supremum is finite, the set of i ∈ I such that ai > 0 is countable. Indeed, for every n ≥ 1, the set is finite, because

If I  is countably infinite and enumerated as I = {i0, i1,...} then the above defined sum satisfies

provided the value ∞ is allowed for the sum of the series.

Any sum over non-negative reals can be understood as the integral of a non-negative function with respect to the counting measure, which accounts for the many similarities between the two constructions.

Abelian topological groups

Let a : IX, where I  is any set and X  is an abelian Hausdorff topological group. Let F  be the collection of all finite subsets of I. Note that F  is a directed set ordered under inclusion with union as join. Define the sum S  of the family a as the limit

if it exists and say that the family a is unconditionally summable. Saying that the sum S  is the limit of finite partial sums means that for every neighborhood V  of 0 in X, there is a finite subset A0 of I  such that

Because F  is not totally ordered, this is not a limit of a sequence of partial sums, but rather of a net.[12][13]

For every W, neighborhood of 0 in X, there is a smaller neighborhood V  such that V − V ⊂ W. It follows that the finite partial sums of an unconditionally summable family ai, i ∈ I, form a Cauchy net, that is: for every W, neighborhood of 0 in X, there is a finite subset A0 of I  such that

When X  is complete, a family a is unconditionally summable in X  if and only if the finite sums satisfy the latter Cauchy net condition. When X  is complete and ai, i ∈ I, is unconditionally summable in X, then for every subset J ⊂ I, the corresponding subfamily aj, j ∈ J, is also unconditionally summable in X.

When the sum of a family of non-negative numbers, in the extended sense defined before, is finite, then it coincides with the sum in the topological group X = R.

If a family a in X  is unconditionally summable, then for every W, neighborhood of 0 in X, there is a finite subset A0 of I  such that ai ∈ W  for every i not in A0. If X  is first-countable, it follows that the set of i ∈ I  such that ai ≠ 0 is countable. This need not be true in a general abelian topological group (see examples below).

Unconditionally convergent series

Suppose that I = N. If a family an, n ∈ N, is unconditionally summable in an abelian Hausdorff topological group X, then the series in the usual sense converges and has the same sum,

By nature, the definition of unconditional summability is insensitive to the order of the summation. When ∑an is unconditionally summable, then the series remains convergent after any permutation σ of the set N of indices, with the same sum,

Conversely, if every permutation of a series ∑an converges, then the series is unconditionally convergent. When X  is complete, then unconditional convergence is also equivalent to the fact that all subseries are convergent; if X  is a Banach space, this is equivalent to say that for every sequence of signs εn = ±1, the series

converges in X. If X  is a Banach space, then one may define the notion of absolute convergence. A series ∑an of vectors in X  converges absolutely if

If a series of vectors in a Banach space converges absolutely then it converges unconditionally, but the converse only holds in finite-dimensional Banach spaces (theorem of Dvoretzky & Rogers (1950)).

Well-ordered sums

Conditionally convergent series can be considered if I is a well-ordered set, for example an ordinal number α0. One may define by transfinite recursion:

and for a limit ordinal α,

if this limit exists. If all limits exist up to α0, then the series converges.

Examples

  1. Given a function f : XY, with Y an abelian topological group, define for every a ∈ X

    a function whose support is a singleton {a}. Then

    in the topology of pointwise convergence (that is, the sum is taken in the infinite product group YX ).
  2. In the definition of partitions of unity, one constructs sums of functions over arbitrary index set I,
    While, formally, this requires a notion of sums of uncountable series, by construction there are, for every given x, only finitely many nonzero terms in the sum, so issues regarding convergence of such sums do not arise. Actually, one usually assumes more: the family of functions is locally finite, i.e., for every x there is a neighborhood of x in which all but a finite number of functions vanish. Any regularity property of the φi,  such as continuity, differentiability, that is preserved under finite sums will be preserved for the sum of any subcollection of this family of functions.
  3. On the first uncountable ordinal ω1 viewed as a topological space in the order topology, the constant function f: [0,ω1) → [0,ω1] given by f(α) = 1 satisfies
    (in other words, ω1 copies of 1 is ω1) only if one takes a limit over all countable partial sums, rather than finite partial sums. This space is not separable.

See also

Notes

  1. ^ Cours d'analyse 1821, p.123 and p.2;   translated into English 2009
  2. ^ No sources are found, presenting a non-contradictory description of such a mathematical notion, positioned in between:   - sequence,  - (real)number,  - expression,  - written symbolic expression,  - mapping of (suitable) sequences on their partial sums sequence, - mapping of (suitable) sequences on the limit of their partial sums sequence.
  3. ^ In his book Calculus (editions 1967 until 2006) Michael Spivak writes:
    The terminology introduced in this definition is usually replaced by less precise expressions; ...
    The statement that {} is, or is not, summable is conventionally replaced by the statement that the series does, or does not, converge.   This terminology is somewhat peculiar, because ...
  4. ^ To simplify wordings, 'sum sequence of' is used to denote the function 'the sequence of partial sums of '.  D.A. Quadling used it in his Mathematical Analysis (editions 1955-1968).
  5. ^ WolframMathWorld:   series expansion,   Maclaurin series]
  6. ^ Cauchy, see p.123 and p.2 quantité C.L.B. Susler, 1828, Susler, S.92, Carl Itzigsohn, 1885, Bradley/Sandifer, 2009
  7. ^ http://arxiv.org/abs/1104.5100/
  8. ^ http://mathworld.wolfram.com/FlintHillsSeries.html
  9. ^ O'Connor, J.J.; Robertson, E.F. (February 1996). "A history of calculus". University of St Andrews. Retrieved 2007-08-07. {{cite web}}: Unknown parameter |lastauthoramp= ignored (|name-list-style= suggested) (help)
  10. ^ Archimedes and Pi-Revisited.
  11. ^ http://www.manchester.ac.uk/discover/news/article/?id=2962
  12. ^ Bourbaki, Nicolas (1998). General Topology: Chapters 1–4. Springer. pp. 261–270. ISBN 9783540642411.
  13. ^ Choquet, Gustave (1966). Topology. Academic Press. pp. 216–231. ISBN 9780121734503.

References

  • Bromwich, T. J. An Introduction to the Theory of Infinite Series MacMillan & Co. 1908, revised 1926, reprinted 1939, 1942, 1949, 1955, 1959, 1965.
  • Dvoretzky, Aryeh; Rogers, C. Ambrose (1950). "Absolute and unconditional convergence in normed linear spaces". Proc. Natl. Acad. Sci. U.S.A. 36 (3): 192–197. doi:10.1073/pnas.36.3.192.
  • Swokowski, Earl W. (1983), Calculus with analytic geometry (Alternate ed.), Boston: Prindle, Weber & Schmidt, ISBN 0-87150-341-7

MR0033975